Photoelectron spectra of hydrazines. IV. Empirical estimation of lone

Oct 1, 1974 - Empirical estimation of lone pair-lone pair dihedral angles and prediction of lone pair ionization potentials for some cyclic and bicycl...
0 downloads 0 Views 705KB Size
6982

of the transition-state protons (fractionation factors The observation of general-base catalysis for the reaction under study5 suggests a transition state of structure 2. Transition-state contributions to the

+iT).

CH;

/

0 ' : HOCH, 2

solvent isotope effect would then come from the bridging proton (fractionation factor +*) and the two solvent protons in the solvation shell of the catalytic methoxide ion (fractionation factors &) so that eq 4 results. k,(l - n

+ 0.74n)3 = kdl - n

+ n$*)(l

-n

+ n&J2

(4)

The latter is easily rearranged to eq 5, of which a plot k,(l

- n + 0 . 7 4 ~ ) ~ / (-1 n + n+m)z = ka(1 - n

+ n+*)

(5)

of the left-hand side us. n should be linear for the correct choice of &. Plots are shown in Figure 1 for +m = 0.8, 1.0, and 1.2. The data are not capable of discriminating strongly among these choices for either substrate. The permitted range of fractionation factors for the bridging proton (determined by the slopes in Figure 1) is +* = 0.25-0.50, corresponding to kH/kD = 2-4. The lower part of this range corresponds to

solvation-bridge effects (2.2 f 0.7),9 and the data are therefore consistent with the view that catalysis in these reactions occurs by the one-proton solvation-bridge mechanism. That is, the formation of the siliconoxygen bond may be assisted by strong hydrogen bonding from the nucleophilic methanol t o a catalyzing base. By analogy with other reaction^,^,^^ this is probably the most reasonable hypothesis, but the data d o not exclude somewhat larger isotope effects for the bridging proton. This could still be consistent with a solvation bridge (which might have a different character and thus a larger isotope effect near a second-row center like silicon), although it might also indicate some participation of the bridging proton in the reaction coordinate.

Experimental Section Materials. Solutions in CH,OH and CHaOD were prepared and manipulated as before. l , l 3 p-Chlorophenoxytriphenylsilane was synthesized following a procedure similar to that which Gerrard and K i l b ~ r n used ' ~ to prepare m-trifluoromethylphenoxytriphenylsilane. Recrystallization from diethyl ether gave the silane in 8 7 % yield, mp 96-98". Anal. Calcd for C21H1qSiOC1:C, 74.50; H, 4.95. Found: C, 74.09; H, 5.09. p-Methoxyphenoxytriphenylsilane was obtained from Schowen6 Kinetic Procedure. Concentrated stock solutions of silanes in dioxane were prepared. Ten microliters of the silane solutions was transferred into the thermostatted cuvet of a Beckman DBG spectrophotometer, and then 2.6 ml of base solutions containing added LiClO, (0.1500 M ) was injected rapidly. Wavelengths used for following the kinetic progress of reaction were 284 nm for p chlorophenoxytriphenylsilaneand 290 nm for p-methoxyphenoxytriphenylsilane. (14) W. Gerrard and K. D. Kilburn, J . Chem. Soc., 1536 (1956).

Photoelectron Spectra of Hydrazines. IV. Empirical Estimation of Lone Pair-Lone Pair Dihedral Angles and Prediction of Lone Pair Ionization Potentials for Some Cyclic and Bicyclic Hydrazines S. F. Nelsen* and J. M. Buschek

Contribution f r o m the Department of Chemisrry, University of Wisconsin, Madison, Wisconsin 53706. Received December 26, 1973 Abstract: The curve calculated for the difference in lone pair ionization potentials (A) as a function of the lone pair-lone pair dihedral angle (0) using INDO calculations on idealized tetramethylhydrazine is presented, and scaling to give more realistic 0 angles from experimental spectra is discussed. The changes in A(0) expected for flattening of the nitrogens from tetrahedral geometry have been calculated. The spectra of five cyclic dialkylhydrazines are given, from which empirical parameters are derived which are used to calculate the average ionization POtentials for 12 cyclic tetraalkylhydrazines. These calculations were fairly successful for cyclic hydrazines, showing A values less than 1 eV, but not satisfactory for tetraalkylhydrazines having large A values. ademacher' and we2 have noted that, since the amount of interaction between the lone pair electrons of a hydrazine should be dependent upon the dihedral anale - 8 between the lone uair orbital axes,

R

(1) (a) P. Rademacher, Angew. Chem., 85, 410 (1973); (b) P. Rademacher, TetrahedronLett., 83 (1974). (2) (a) S . F. Nelsen and J. M. Buschek, J . Amer. Chem. Soc., 95, 201 1 (1973); (b) ibid.,95,2013 (1973); (c) ibid., 96,2392 (1974).

Journal of the American Chemical Society J 96:22

photoelectron spectroscopy (pes) should be of use in determination of hydrazine conformations. As Hoffmann3 has pointed out, the lone pair orbitals will interact to form two molecular orbitals, of which the symmetric lone pair combination (n+) will be bonding and hence lower in energy (higher in ionization poten(3) R. Hoffmann, Accounts Chem. Res., 4 , l (1971).

October 30, 1974

6983

n-

n+

tial) at 6' values below about 90", whereas the antisymmetric (n-) combination will be lowest in energy at high 8 values. The lone pair ionizations of saturated aliphatic hydrazines are usually well separated from ubond ionizations, allowing measurement of their positions. We choose to express these ionizations in terms of their average, IP,, = (IPl IP2)/2, and their separation, A = IP2 - IP1, where IP1 is the lowest ionization potential. For the series of the five methylated hydrazines, we found that calculated orbital energies (INDO approximation, using idealized geometries and 6' = 90" for all five compounds), Ei,when plotted against the ionization potentials gave a good straight line having a standard deviation of 42 meV for the ten points.2c Since our experimental error is (conservatively) 30 meV, it appears that the INDO calculations adequately correlate both A and IP,,, suggesting that the methylated hydrazines have roughly similar 6' values. Calculations on hydrazine and tetramethylhydrazine in which 6' was varied from 0 t o 180", invoking Koopman's theorem t o equate - E with IP, give the prediction that A follows a rough cos 6' relationship, although the magnitudes of A(0") and A(l80") were different, and the "crossover points" (where n+ and n- have the same energy, A = 0) were slightly different (83" for tetramethylhydrazine, 85.5" for hydrazine), and the tetramethylhydrazine curve was rather flatter than that of hydrazine near 180". This suggests that a A(@ function, t o be used as a working curve for evaluation of 6' from experimental A values, ought to take into account the substitution pattern of the hydrazine ; hydrazine itself is expected t o show a larger A value for a given 0 value than is tetramethylhydrazine (except very near the crossover point). Since it is impractical to perform MO calculations on all molecules for which we would like to interpret pes spectra, we have made the assumption that substitution of one alkyl group for another (if 6' does not change) is a small effect which will show up principally in IP,,, leaving A unchanged. The independence of A with substituent changes was borne out experimentally, since nine tetraalkylhydrazines with Me, Et, i-Pr, n-Pr, n-Bu, and t-Bu substituents showed quite constant A values 2c (eight in the range 0.51-0.55 eV, and only the most hindered one, triisopropylmethylhydrazine, having a slightly larger splitting of 0.61 eV). The IP,, values ranged from 8.55 t o 7.90 for this series. A quantitative method of accounting for the IP,, values was devised, defining a parameter, X(R) (eq l), obtained from monoalkylhydrazine pes IP,, values, and accounting for the multiple substitution of tetraalkyhydrazines using eq 2,2c where (LS,90") refers to least-square values from

+

X(R) IP,(calcd)

=

IP,,(RNHNH2)/9.872

(1)

IPa,(LS,90")~(X(R)f A(LS,90")/2)

(2)

=

R

the observed us. calculated methylated hydrazine plot and the plus sign gives IP2. Using eq 1 and 2, a series

of 22 acyclic alkylated hydrazines gave an excellent straight line in an IPi(obsd) us. IPi(calcd) plot, which had a standard deviation of 78 meV (44 points). Our procedure is empirically justified by the remarkable success of eq 1 and 2 in predicting the pes spectra of acyclic hydrazines. Improved fit could certainly be obtained by considering the slightly different 6' values which are present for hydrazines of different substitution patterns, but this imposes a degree of confidence in knowing the crossover points which is probably greater than is presently justified. In this paper we consider some aspects of the use of pes for obtaining 6' values of hydrazines.

Results and Discussion Angle Dependence of Tetraalkylhydrazine Pes. The substantial substituent independence of acyclic hydrazine pes spectra certainly does not extend to cyclic compounds.'I2 Since most of our work has been with tetraalkylhydrazines, we shall restrict our discussion to these compounds. The INDO predictions for the splittings of tetramethylhydrazine (idealized geometry) as a function of 6' are shown in Table I as AI(@ (the I is Table I. Angle-Dependent Terms for Correlation of Ionization Potentials and 8 for Tetraalkylhydrazines

0 10 30 60 80 90 100 120 140 160 170 180

AdO): eV

Ap.d8)? eV

U(0); eV

3.083 2,944 2.231 0.999 0.133 0.354 0,863 1.823 2.585 3.184 3.423 3.548

2.172 2.074 1.572 0.704 0.094 0.249 0.608 1.284 1.821 2.244 2.412 2.500

0.275 0.249 0.154 0.035 0.011 0.004 0 0.045 0.214 0.429 0.527 0.581

a AI(@ = El - E2, where El is the highest occupied MO (n- for 0 < 83", n+ for 8 > 83") and Enthe other lone pair MO, from INDO calculations on tetramethylhydrazine in standard geometry. b A,.,(O) = 0.705AI(0), used as a working curve for evaluation of 0 from observed A. c U(0) = E,d100") - E,(@.

for INDO). Not surprisingly, direct use of AI(@ for prediction of 6' gives unreasonable values. INDO overestimates the splitting, and AI(@ must be scaled down to be useful. We have chosen t o simply multiply A,(@ by a constant t o perform this scaling. Consideration of all of our data suggests that A(180") is about 2.5 eV, so our working curve for estimation of angles from A values was obtained by multiplying Ado) by 0.705, giving the A2.5(6')values listed in Table 1.j (4) The crossover point becomes crucial in fine adjustments of 0, and, for low A values, it is not clear whether one is on the high or low 8 side of the crossover point. ( 5 ) Rademacherlb has suggested using the notably simpler function for what we call A (he uses a sign change depending on whether n, or n is higher in energy), A E = 2.17 cos 0 - 0.35. Our scaling only differs significantly from his near 0 = 0". AE = 0 at 80.7", compared to 83" for A 2 . 6 . On the low 0 side A 2 . S and AE differ by a maximum of 3" until A is >1.57 eV, and, on the high 0 side, the use Of A2.5 will lead to higher 0 values by an amount which increases from 2.3" (A = 0) to 5' at A = 1.37, passes through a maximum of 9 " at 2.25 eV, and then decreases (AE(180") = (-)2.52). The similarity of these functions suggests that fair accuracy might be expected for the derived 0 values, since the methods of obtaining A E and A2.s were rather different.

Nelsen, Buschek

Photoelectron Spectra of Hydrazines

6984

The INDO calculations also show that E,, depends upon 6, with an extremum near 100" (see Table I). Taking this into account, we examine using equations 3 and 4 for prediction of ionization potentiaIs of tetra-

]Pi(@

=

IPav(0) *

&,,(e)

(3) (4)

alkylhydrazines. It should be noted that SI(0) is a positive number which is negligible (